Determining Number of Active Sites and TOF for ... - ACS Publications

Feb 11, 2016 - ... with C16O2/C18O2 isotope switch and H2-TPR experiments, for the first time that (i) the high-temperature water–gas shift (HT-WGS)...
0 downloads 0 Views 642KB Size
Letter pubs.acs.org/acscatalysis

Determining Number of Active Sites and TOF for the HighTemperature Water Gas Shift Reaction by Iron Oxide-Based Catalysts Minghui Zhu and Israel E. Wachs* Operando Molecular Spectroscopy & Catalysis Laboratory, Department of Chemical and Biomolecular Engineering, Lehigh University, Bethlehem, Pennsylvania 18015, United States S Supporting Information *

ABSTRACT: This study demonstrates, with C16O2/C18O2 isotope switch and H2-TPR experiments, for the f irst time that (i) the high-temperature water−gas shift (HT-WGS) reaction by copper−chromium-iron oxide catalysts follows a redox mechanism dominated by the surface layer, (ii) the number of catalytic active sites can be quantified by the isotopic switch, and (iii) the turnover frequency (TOF) can be determined from knowledge of the number of sites. The quantitative TOF values reveal that chromium is only a textural promoter, whereas copper is a chemical promoter. KEYWORDS: water−gas shift, iron-based, isotope switch, promoters, number of active sites, turnover frequency, surface oxygen

I

Fe2+ being oxidized to bulk Fe3+ by H2O and bulk Fe3+ becoming reduced to bulk Fe2+ by CO.10,11 While numerous detailed redox mechanisms were proposed, 12−20 direct experimental proof is still lacking and no information has been provided about the catalyst surface during HT-WGS reaction conditions. During the HT-WGS reaction, the equilibrated bulk iron oxide phase is present as magnetite (Fe3O4), which is produced by the partial reduction of the starting hematite (Fe2O3) phase.21,22 Chromium oxide is added as a textural promoter to inhibit sintering and stabilize the surface area of the magnetite. The chemical promotion of magnetite by chromia during the HTS reaction has been proposed and several models have been given.12−14,23,24 Copper is also added as a promoter in commercial iron−chromium oxide catalyst to increase the activity over a wider temperature range.25,26 The promotion mechanism of copper during HT-WGS has received extensive discussion without reaching a consensus.23,25−31 Recently, there is much interest in developing a Cr-free iron oxide HT-WGS catalyst because of the presence of toxic chromium(VI) oxide in this catalyst.32 The lack of fundamental understanding of how the HT-WGS catalyst functions, however, hampers the development of Cr-free HT-WGS iron oxide-based catalysts. This work provides direct experimental evidence about fundamental aspects of the HT-WGS reaction by Fe-based catalysts: (i) reaction mechanism of the HT-WGS catalytic reaction by the Cr2O3−Fe2O3 mixed oxide catalyst (ii) number of catalytic active sites, (iii) nature of the most abundant reaction intermediate (mari), (iv)specific reaction rates (TOF =

ndustrial H2 is currently primarily produced by methane steaming reforming (MSR) followed by the water−gas shift (WGS) reaction to increase or control the H2/CO ratio and is employed in numerous applications like ammonia synthesis (from H2/N2), methanol synthesis (from H2/CO/CO2), synthetic fuels (from H2/CO), and so on. The WGS reaction involves reaction of carbon monoxide with steam to produce H2 and CO2 and is commercially performed in several temperature stages with different catalysts to optimize the greater CO equilibrium conversion attained at lower temperatures because the reaction is exothermic and reversible.1,2 CO + H 2O ↔ CO2 + H 2

ΔH = − 40.6kJ/mol

(1)

The high-temperature water−gas shift (HT-WGS) reaction is commercially performed at ∼350−450 °C with iron-based catalysts and the low temperature water−gas shift (LT-WGS) reaction is performed at ∼190−250 °C with copper-based catalysts. The reaction mechanism of the HT-WGS reaction catalyzed by iron−chromium oxide based catalyst has been extensively studied without reaching general agreement.3 Armstrong and Hilditch were the first to propose a mechanism that involves a surface reaction intermediate such as surface formate (HCOO*) which is referred to as the associative mechanism.4 Subsequent experimental and modeling studies of the hightemperature WGS reaction have been inconclusive, both supporting and contradicting the presence of a surface formate intermediate.5,6 The most accepted mechanism, however, is the “regenerative” or redox mechanism involving alternate reduction of the oxidized catalyst by CO and oxidation of the reduced catalyst by H2O.2,5,7−9 The importance of the redox mechanism for this HT-WGS catalyst has been confirmed by the observation of the bulk Fe2+/Fe3+ redox couple, with bulk © XXXX American Chemical Society

Received: December 28, 2015 Revised: February 7, 2016

1764

DOI: 10.1021/acscatal.5b02961 ACS Catal. 2016, 6, 1764−1767

Letter

ACS Catalysis turnover frequency), and (v) promotion mechanisms of Cr and Cu. The specific catalytic activity allows for the first time quantitative comparison of HT-WGS iron oxide-based catalysts and determining the promotion of Cr and Cu. Such fundamental information establishes the foundation for the rational design of Cr-free iron-based HT-WGS catalysts.



REACTION MECHANISM The steady-state isotope switch experiment was performed with the iron−chromium oxide catalyst, and the time-resolved MS signals are presented in Figure 1 (the experimental details are

Figure 2. H2-TPR profile of Cr2O3−Fe2O3 catalyst after the steadystate isotope switch experiment. Catalyst was cooled in flowing He and then heated at 10 °C/min of flowing 10% H2/He (30 mL/min).

to the surface region. By integrating the H2-TPR isotopic water peaks, only ∼8% of the total oxygen in the equilibrated catalyst is involved in the steady-state isotope switch experiment. These isotopic oxygen exchange studies prove for the f irst time that the HT-WGS reaction by chromium−iron oxide catalysts follows a redox reaction mechanism and not an associative reaction mechanism involving a surface reaction intermediate (e.g., surface HCOO*). The redox process is dominated by a surface Mars−van Krevelen (MVK) reaction mechanism, where only the catalyst surface layer is rapidly exchanging oxygen with the reactants, and the catalyst bulk lattice MVK mechanism also contributes to the oxygen exchange by slower diffusion over an extended period of reaction time.

Figure 1. Transient response of H2, C16O2, C16O18O, C18O2, H216O, and H218O during steady-state isotope switch from C16O2 + H2 to C18O2 + H2 on Cr2O3−Fe2O3 catalyst (T = 400 °C).



given in the Supporting Information section). Upon isotope switch (C16O2/H2 → C18O2/H2), the H2 signal remains constant while the C16O2 signal sharply decreases and the C18O2 signal increases. The increase in C18O2 is slightly slower than the decrease in C16O2 because of the transient production of C16O18O during the isotope switch. The production of C16O18O also shows that oxygen exchange is taking place between the reactants and the oxygen from the catalyst. The decrease of the H216O signal is slightly slower than the decrease of the C16O2 signal reflecting the longer holdup of moisture than carbon dioxide in the catalyst bed. The entire isotope switch response takes place in ∼2 min, which demonstrates that only a finite amount of oxygen is involved. To gain insight into the oxygen isotopes remaining in the catalyst after the isotope switch, H2-TPR was performed afterward to monitor the population of 18O and 16O in the catalyst by formation of the corresponding isotopic water and presented in Figure 2. The production of water between ∼200−300 °C corresponds to the reduction of surface oxygen from the catalyst and yields comparable amounts of H218O/H216O ∼ 1. The presence of the doublet in the H2-TPR spectra suggests that two distinct oxygen sites may be participating in the reduction process, but the identity of the participating oxygen sites are not known. The production of water above 350 °C corresponds to the reduction of bulk lattice oxygen from the catalyst and the H218O/H216O ≪ 1. Some surface 16O and bulk 18O was observed, revealing that oxygen exchange is also taking place between the surface and bulk phases, which may be facilitated by the reduction process, but the exchange is mostly confined

MOST ABUNDANT REACTIVE INTERMEDIATES (MARI) AND NUMBER OF ACTIVE SITES (NS) The total oxygen participating in the HT-WGS redox process with the chromium−iron oxide catalyst was quantified by the isotope switch after inert flush experiment (C16O2/H2 → He → C18O2/H2), and the time-resolved evolution of the products is shown in Figure 3. The isotope switch experiment was performed at 330 °C because this reaction temperature provides differential reaction conditions (conversions Cr2O3−Fe2O3 > Fe2O3) with Cr enhancing the reaction rate per gram by ∼2× and Cu by an additional ∼3×. The double Cr−Cu promoted iron oxide catalyst is ∼5× more active than unpromoted iron oxide per gram of catalyst. The corresponding TOF values (TOF = activity/Ns) indicate that the specific TOF value for HT-WGS by the chromium−iron oxide catalyst is essentially the same as the unpromoted iron oxide catalyst. Thus, Cr is a textural promoter that increases the number of catalytic active sites by stabilizing higher surface area iron oxide, but Cr does not chemically promote the HT-WGS reaction by iron oxide. In contrast, Cu is a chemical promoter increasing the TOF value by ∼3× compared to the TOF values for Cr 2 O 3 −Fe 2 O 3 and unpromoted Fe2O3 catalysts. In conclusion, the HT-WGS reaction by iron-based catalysts follows a redox mechanism primarily involving oxygen atoms from the surface layer, and the participating oxygen atoms represent the most abundant reactive intermediate (mari). The isotopic switch experiments allow for the f irst time determination of the number of catalytic active sites and specific catalytic reactivity (TOF). The Cr is a textural promoter that increases the number of participating oxygen sites by stabilizing iron oxides with higher surface areas. The Cu is a chemical promoter that increases the specific reaction rate (TOF) of the HT-WGS reaction by iron-based catalysts. The dual promotion of iron oxide by Cr and Cu yields a HT-WGS catalyst that has a specific reaction rate (TOF) that is ∼3× greater and a catalyst activity per gram that is ∼5× greater than an unpromoted iron oxide catalyst.

Figure 3. Transient response of He, H2, C16O2, C16O18O, C18O2, C16O, C18O, H216O, and H218O during isotope switch after inert flush on Cr2O3−Fe2O3 (T = 330 °C). The MS signals for all products were normalized to the same maximum and minimum intensity for better comparison of their transient behavior. The CO isotope signals are corrected for contribution of CO2 cracking in the MS because cracking of the dominant CO2 isotopes in the mass spectrometer significantly contribute to the CO MS signals.

the oxygen isotope population on the catalyst surface during the HT-WGS reaction. The first water isotope to form was H216O followed by appearance of H218O. The appearance of the water isotopes (H216O/H218O) lags the corresponding carbon dioxide isotopes (C16O2/C18O2 and C16O/C18O), respectively, reflecting the longer holdup of water than carbon oxides in the catalyst bed and walls of the capillary to the MS. The longer evolution of H216O is mostly related to the holdup of moisture in the reactor system (catalyst and walls of the capillary to the MS) and possibly also a second slower oxygen exchange process related to slow diffusion of bulk lattice oxygen to the surface of the catalyst. The isotopic switch findings also demonstrate that the most abundant reactive intermediate (mari) for the HT-WGS reaction by chromium−iron oxide catalysts is the reactive O*. The isotope switch experiment after the inert flush also provides for quantification of the number of oxygen sites participating in the HT-reverse WGS, as well as HT-forward WGS because of well-known concept of microscopic reversibility,33 over the chromium−iron oxide catalyst by counting the number of 16O atoms in the reaction products (C16O2, C16O18O, C16O, and H216O). The number of oxygen sites per gram of each catalyst is given in Table 1. Both Crpromoted catalysts have almost twice as many active sites per gram as the unprompted iron oxide catalyst that is primarily related to the higher surface area of the Cr-promoted catalysts (Table S2). The density of the active sites was calculated by dividing the Ns by specific surface area of activated catalysts

Table 1. WGS Activity, Number of Sites [n(16O)], and Turnover Frequencies (TOFs)a

a

catalyst

WGS activity-H2O conversion (× 10−6 mol/s·g)

Ns: n(16O) (× 10−3 mol/g)

density of Ns (16O atoms/nm2)

TOF (× 10−3 s−1)

Fe2O3 Cr2O3−Fe2O3 CuO-Cr2O3−Fe2O3

1.2 ± 0.1 2.0 ± 0.1 5.7 ± 0.3

0.9 ± 0.1 1.7 ± 0.2 1.6 ± 0.2

19 ± 2 16 ± 2 16 ± 2

1.3 ± 0.2 1.2 ± 0.2 3.5 ± 0.4

10% CO/Ar (10 mL/min), He (30 mL/min), and water vapor (H2O/CO ∼ 1); T = 330°C. 1766

DOI: 10.1021/acscatal.5b02961 ACS Catal. 2016, 6, 1764−1767

Letter

ACS Catalysis



(25) Andreev, A.; Idakiev, V.; Mihajlova, D.; Shopov, D. Appl. Catal. 1986, 22, 385−387. (26) Idakiev, V.; Mihajlova, D.; Kunev, B.; Andreev, A. React. Kinet. Catal. Lett. 1987, 33, 119−124. (27) Grunwaldt, J. D.; Kappen, P.; Hammershoi, B. S.; Troger, L.; Clausen, B. S. J. Synchrotron Radiat. 2001, 8, 572−574. (28) Kappen, P.; Grunwaldt, J. D.; Hammershoi, B. S.; Troger, L.; Clausen, B. S. J. Catal. 2001, 198, 56−65. (29) Rhodes, C.; Williams, B. P.; King, F.; Hutchings, G. J. Catal. Commun. 2002, 3, 381−384. (30) Estrella, M.; Barrio, L.; Zhou, G.; Wang, X. Q.; Wang, Q.; Wen, W.; Hanson, J. C.; Frenkel, A. I.; Rodriguez, J. A. J. Phys. Chem. C 2009, 113, 14411−14417. (31) Puig-Molina, A.; Cano, F. M.; Janssens, T. V. W. J. Phys. Chem. C 2010, 114, 15410−15416. (32) Lee, D. W.; Lee, M. S.; Lee, J. Y.; Kim, S.; Eom, H. J.; Moon, D. J.; Lee, K. Y. Catal. Today 2013, 210, 2−9. (33) Boudart, M.; Djéga-Mariadassou, G. Kinetics of Heterogeneous Catalytic Reactions; Princeton University Press: Princeton, NJ, 1984. (34) Sharp, J. C.; Yao, Y. X.; Campbell, C. T. J. Phys. Chem. C 2013, 117, 24932−24936. (35) Joseph, Y.; Kuhrs, C.; Ranke, W.; Ritter, M.; Weiss, W. Chem. Phys. Lett. 1999, 314, 195−202.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.5b02961. Catalyst synthesis and preparation; isotope switch experimental details; activity measurement details; flow BET surface area measurement details; BET surface areas (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS The authors acknowledge financial support from National Science Foundation Grant CBET - 1511689. REFERENCES

(1) Newsome, D. S. Catal. Rev.: Sci. Eng. 1980, 21, 275−318. (2) Ratnasamy, C.; Wagner, J. P. Catal. Rev.: Sci. Eng. 2009, 51, 325− 440. (3) Smith, R. J. B.; Loganathan, M.; Shantha, M. S. Int. J. Chem. React. Eng. 2010, 8, DOI: 10.2202/1542-6580.2238. (4) Armstrong, E. F.; Hilditch, T. P. Proc. R. Soc. London, Ser. A 1920, 97, 265−273. (5) Boudjemaa, A.; Daniel, C.; Mirodatos, C.; Trari, M.; Auroux, A.; Bouarab, R. C. R. Chim. 2011, 14, 534−538. (6) Diagne, C.; Vos, P. J.; Kiennemann, A.; Perrez, M. J.; Portela, M. F. React. Kinet. Catal. Lett. 1990, 42, 25−31. (7) Rhodes, C.; Hutchings, G. J.; Ward, A. M. Catal. Today 1995, 23, 43−58. (8) Ladebeck, J. R.; Wagner, J. P. In Handbook of Fuel Cells; Wolf, W., Lamm, A., Gasteiger, H. A., Eds.; John Wiley & Sons, Ltd: Chichester, U.K., 2003; pp 190−201. (9) Kubsh, J. E.; Dumesic, J. A. AIChE J. 1982, 28, 793−800. (10) Boreskov, G. K.; Yurieva, T. M.; Sergeeva, A. S. Kinet. Catal. 1970, 11, 374−381. (11) Khan, A.; Chen, P.; Boolchand, P.; Smirniotis, P. G. J. Catal. 2008, 253, 91−104. (12) Chinchen, G. C.; Logan, R. H.; Spencer, M. S. Appl. Catal. 1984, 12, 89−96. (13) Chinchen, G. C.; Logan, R. H.; Spencer, M. S. Appl. Catal. 1984, 12, 69−88. (14) Chinchen, G. C.; Logan, R. H.; Spencer, M. S. Appl. Catal. 1984, 12, 97−103. (15) Keiski, R. L.; Salmi, T.; Niemistö, P.; Ainassaari, J.; Pohjola, V. J. Appl. Catal., A 1996, 137, 349−370. (16) Salmi, T.; Lindfors, L. E.; Bostrom, S. Chem. Eng. Sci. 1986, 41, 929−936. (17) Oki, S.; Mezaki, R. Ind. Eng. Chem. Res. 1988, 27, 15−21. (18) Mezaki, R.; Oki, S. J. Catal. 1973, 30, 488−489. (19) Tinkle, M.; Dumesic, J. A. J. Catal. 1987, 103, 65−78. (20) Tinkle, M.; Dumesic, J. A. J. Phys. Chem. 1984, 88, 4127−4130. (21) Kundu, M. L.; Sengupta, A. C.; Maiti, G. C.; Sen, B.; Ghosh, S. K.; Kuznetsov, V. I.; Kustova, G. N.; Yurchenko, E. N. J. Catal. 1988, 112, 375−383. (22) Patlolla, A.; Carino, E. V.; Ehrlich, S. N.; Stavitski, E.; Frenkel, A. I. ACS Catal. 2012, 2, 2216−2223. (23) Edwards, M. A.; Whittle, D. M.; Rhodes, C.; Ward, A. M.; Rohan, D.; Shannon, M. D.; Hutchings, G. J.; Kiely, C. J. Phys. Chem. Chem. Phys. 2002, 4, 3902−3908. (24) Natesakhawat, S.; Wang, X. Q.; Zhang, L. Z.; Ozkan, U. S. J. Mol. Catal. A: Chem. 2006, 260, 82−94. 1767

DOI: 10.1021/acscatal.5b02961 ACS Catal. 2016, 6, 1764−1767